Sunteți pe pagina 1din 18

Journal of South American Earth Sciences 32 (2011) 246e263

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

Depositional and provenance record of the Paleogene transition from foreland


to hinterland basin evolution during Andean orogenesis, northern Middle
Magdalena Valley Basin, Colombia
Christopher J. Moreno a,1, Brian K. Horton a, b, *, Victor Caballero c, d, Andrs Mora d,
Mauricio Parra a, d, Jair Sierra d
a

Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, TX 78712, USA
Institute for Geophysics, Jackson School of Geosciences, University of Texas at Austin, Austin, TX 78712, USA
Escuela de Geologa, Universidad Industrial de Santander, Bucaramanga, Colombia
d
Instituto Colombiano del Petrleo, Ecopetrol, Bucaramanga, Colombia
b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 7 September 2010
Accepted 29 March 2011

The Central Cordillera and Eastern Cordillera of the northern Andes form the topographic anks of the
north-trending Magdalena Valley Basin. Constraining the growth of these ranges and intervening basin
has implications for Andean shortening and the transformation from a foreland to hinterland basin
conguration. We present sedimentological, paleocurrent, and sandstone petrographic results from
Cenozoic type localities to provide insights into the tectonic history of the northern Middle Magdalena
Valley Basin of Colombia. In the Nuevo Mundo Syncline, the mid-Paleocene transition from marine to
nonmarine deposystems of the Lisama Formation corresponds with a paleocurrent shift from northward
to eastward transport. These changes match detrital geochronological evidence for a contemporaneous
shift from cratonic (Amazonian) to orogenic (Andean) provenance, suggesting initial shortening-related
uplift of the Central Cordillera and foreland basin generation in the Magdalena Valley by mid-Paleocene
time. Subsequent establishment of a meandering uvial system is recorded in loweremiddle Eocene
strata of the lower La Paz Formation.
Eastward paleocurrents in mid-Paleocene through uppermost Eocene uvial deposits indicate
a continuous inuence of western sediment source areas. However, at the upper middle Eocene
(w40 Ma) boundary between the lower and upper La Paz Formation, sandstone compositions show
a drastic decrease in lithic content, particularly lithic volcanic fragments. This change is accompanied by
a facies shift from mixed channel and overbank facies to thick, amalgamated braided uvial deposits of
possible uvial megafans, reecting changes in both the composition and proximity of western sediment
sources. We attribute these modications to the growing inuence of exhumed La Cira-Infantas paleohighs in the axial Magdalena Valley, features presently buried beneath upper EoceneeQuaternary basin
ll along the western ank of the Nuevo Mundo Syncline.
In uppermost Eocene strata of the lower Esmeraldas Formation, paleocurrents show a sharp reversal
from eastward to dominantly westward transport that persisted into the Neogene. The Esmeraldas also
records a change to more-distal, oodplain-dominated deposition of ner sediments. These adjustments
are interpreted to reect burial of the La Cira-Infantas highs and onset of Eastern Cordillera exhumation,
resulting in a transition from foreland to hinterland basin conditions in the Magdalena Valley. The lack of
signicant variation in sandstone compositions suggests a bulk-rock compositional similarity between
the La Cira-Infantas paleohighs (subsurface Magdalena Valley) and the Eastern Cordillera. Collectively,
the data presented here rene previous thermochronologic and provenance studies and suggest that
major uplift-induced exhumation in the Central Cordillera and Eastern Cordillera commenced by the
mid-Paleocene and latest Eocene, respectively.
2011 Elsevier Ltd. All rights reserved.

Keywords:
Andes
Colombia
Eastern Cordillera
Llanos basin
Magdalena valley
Fold-thrust belts
Foreland basins
Provenance
Sedimentary petrology
Stratigraphy

* Corresponding author. Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, Texas 78712, USA. Tel.: 1 512 471 1869.
E-mail address: horton@mail.utexas.edu (B.K. Horton).
1
Present address: Mack Energy Corporation, 201 Main Street, Suite 1660, Fort Worth, Texas 76102, USA.
0895-9811/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsames.2011.03.018

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

1. Introduction
Although South American foreland basins represent clear
signals of Andean orogenesis (Jordan, 1995; Cooper et al., 1995;
Horton and DeCelles, 1997; Bayona et al., 2008), hinterland basins
situated within modern intermontane valleys and plateau regions
of the Andes constitute protracted, nearly continuous records of
Cenozoic synorogenic sedimentation (Van Houten and Travis, 1968;
Jordan and Alonso, 1987; Marocco et al., 1995; Horton et al., 2002;
Horton, 2005, in press; Leier et al., 2010; Murray et al., 2010).
Among the best-known cases is the Magdalena Valley Basin in the
northern Andes of Colombia, where siliciclastic basin ll up to
10 km thick is located in an intermontane setting between the
Central and Eastern Cordilleras (Fig. 1). A long history of hydrocarbon exploration combined with sufcient surface and subsurface control make the Magdalena Valley an important province for
understanding deformation and basin evolution in an inter-Andean
setting. Within the basin, the Nuevo Mundo Syncline (Fig. 2)
contains critical type localities for many Cenozoic stratigraphic
units identied in the Middle Magdalena Valley Basin (Pilsbry and
Olsson, 1935; Morales et al., 1958; Ramirez, 1988; Schamel, 1991;
Surez, 1997; Ramn, 1998; Gmez, 2001; Gmez et al., 2003,
2005a, 2005b; Pardo-Trujillo, 2004; Rolon, 2004; Nie et al., 2010).
Understanding basin evolution in the Magdalena Valley represents a key component in evaluating the tectonic evolution of the
northern Andes. Unresolved issues include the competing inuences of the Nazca and Caribbean plates (Pennington, 1981; van der
Hilst and Mann, 1994; Kellogg and Vega, 1995; Taboada et al., 2000;
Corts et al., 2005), the degree of structural inheritance from largescale Mesozoic rifting (Cooper et al., 1995; Mora et al., 2006;
Sarmiento-Rojas et al., 2006), and the role of climate and
orographic barriers in the exhumation history of the orogenic belt
(Mora et al., 2008; Horton et al., 2010a). Additional targeted questions for the Magdalena Valley relate to the complex pattern of
basin lling and varied structural styles. For example, stratigraphic

247

studies have shown difculty in regional correlation among


different parts of the basin (Ramon and Rosero, 2006; Rincn et al.,
2007), often due to rapid lateral facies variations involving multiple
conglomeratic and volcaniclastic levels (Morales et al., 1958; Van
Houten, 1976). In addition, surface and subsurface data reveal
a combination of thin- and thick-skinned, east- and west-vergent
structures (Julivert, 1970; Butler and Schamel, 1988; Dengo and
Covey, 1993; Cooper et al., 1995; Restrepo-Pace et al., 2004), kinematic histories of fault reactivation and inversion (Gmez et al.,
2005a; Ramon and Rosero, 2006), and potentially signicant
components of strike-slip displacement (Montes et al., 2005;
Acosta et al., 2007).
This paper evaluates the sedimentology and provenance of siliciclastic rocks that record the Cenozoic interactions among different
sediment sources in the Middle Magdalena Valley Basin and adjacent regions. In light of proposed estimates of initial uplift of the
Central Cordillera in latest CretaceouseEocene time and the Eastern
Cordillera in PaleoceneeOligocene time (Villamil, 1999; Gmez
et al., 2003, 2005a, 2005b; Bayona et al., 2008; Parra et al., 2009a,
2009b; Horton et al., 2010b; Mora et al., 2010; Nie et al., 2010;
Saylor et al., 2011), we anticipate that Paleogene strata should
record key temporal shifts in sediment provenance attributable to
the initiation of uplift in these bounding ranges. We further recognize the importance of the presently buried La Cira-Infantas paleohighs (Fig. 2B) as potential sources of sediment during early
evolution of the northern Middle Magdalena Valley Basin. Many
previous studies have focused on stratigraphy, seismic reection,
and well-log data with an emphasis on petroleum reservoir characteristics (Morales et al., 1958; Ramirez, 1988; Surez, 1997), or
have utilized a regional structural perspective (Dengo and Covey,
1993; Cooper et al., 1995) or biostratigraphic framework (Pilsbry
and Olsson, 1935; Hopping, 1967; Pardo-Trujillo, 2004). Although
recent investigations appropriately emphasize the Cenozoic timing
constraints offered by growth stratal records (e.g., Gmez, 2001;
Gmez et al., 2005b), possible limitations include incomplete

Fig. 1. Map of the Colombian Andes depicting regional topography and major tectonic-geomorphic provinces (WCdWestern Cordillera; CVdCauca Valley; CCdCentral Cordillera;
MMVdMiddle Magdalena Valley Basin; ECdEastern Cordillera; SantandereSantander Massif).

248

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

Fig. 2. (A) Geologic map of the Nuevo Mundo Syncline on the eastern ank of the northern Middle Magdalena Valley Basin (after Caballero, 2010; Caballero et al., 2010) (see Fig. 1
for location), showing locations of geologic cross section (Fig. 2B), seismic reection prole (Fig. 3), and measured stratigraphic section (Fig. 4) near the trace of the Sogamoso river.
(B) Approximately WNW-ESE cross section identifying major structural and stratigraphic elements of the northern Middle Magdalena Valley Basin (modied from Pardo-Trujillo,
2004; Gmez et al., 2005b).

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

preservation, poor seismic resolution, and complex fault-fold


mechanisms (Suppe et al., 1992; Allmendinger, 1998). The current
study seeks to assess the Paleogene deposystems in the type locality
of the Middle Magdalena Valley Basin, with an emphasis on variations in sediment dispersal patterns and sediment source areas in
the context of potential evolving structures. Considerations of
provenance for the exposed succession in the Nuevo Mundo
Syncline are further aided by recent detrital zircon UePb geochronological data (Horton et al., 2010a, 2010b; Nie et al., 2010; Saylor
et al., 2011) that help to discriminate sediment contributed from
the western and eastern topographic margins of the Magdalena
Valley.

2. Geologic setting
2.1. Regional tectonic context
The northern Andes of Colombia consist of three major mountain ranges and their corresponding intermontane and foreland
basins (Fig. 1). From west to east, the principal tectonic and
geomorphic features include the Western Cordillera, Cauca Valley,
Central Cordillera, Magdalena Valley, Eastern Cordillera, and Llanos
foreland basin. Farther east, the Guyana Shield constitutes
Precambrian continental basement of the northern Amazonian
Craton (Teixeira et al., 1989; Cordani et al., 2000). The Western
Cordillera originated as an allochthonous oceanic terrane accreted
at w65e60 Ma (Aspden and McCourt, 1986). The Romeral Fault
largely following the Cauca Valley between the Western and
Central Cordilleras (Fig. 1) forms the boundary between South
American Precambrian basement and accreted oceanic terranes
(Gmez et al., 2003). The Western Cordillera comprises oceanic
basalt, tuff, and volcaniclastic sedimentary and metasedimentary
rock of dominantly Late Cretaceous age overprinted by Cenozoic
igneous intrusions (McCourt et al., 1984; Aspden and McCourt,
1986; Aspden et al., 1987). Because early uplift of the Central
Cordillera formed a topographic barrier, the Western Cordillera is
not an important sediment source to the Magdalena Valley (Gmez
et al., 2003).
Central Cordillera uplift began in a northward-propagating
pattern as a result of the subduction-related compression and
accretion of the Western Cordillera (Gmez et al., 2003, 2005a,
2005b). The Central Cordillera consists of Mesozoic arc-related
rocks (principally Jurassic granodiorite/tonalite and Cretaceous
diorite/quartz diorite, but also Lower Cretaceous basalt) developed
on a belt of Paleozoic metamorphic rock (McCourt et al., 1984;
Aspden et al., 1987). Jurassic-Cretaceous pyroclastic deposits are
also reported for the San Lucas range (Kammer and Snchez, 2006),
which forms the northeastern margin of the Central Cordillera
(Fig. 1).
The Eastern Cordillera is an asymmetric, doubly vergent
fold-thrust belt formed by Cenozoic east-west compression and
associated inversion of Mesozoic normal faults (Corredor, 2003;
Gmez et al., 2005b; Mora et al., 2006). Uplift of the Eastern
Cordillera partitioned the Cretaceous-Paleogene retroarc basin,
resulting in the generation of the intermontane Magdalena Valley
Basin on the western ank and the foreland Llanos Basin on the
eastern ank (Fig. 1). Precambrian-Paleozoic metamorphic and
igneous basement of the Eastern Cordillera is overlain by PaleozoiceMesozoic sedimentary rocks, which are dominated volumetrically by Cretaceous marine clastic rocks, but also contain
some red beds, volcanic rocks, and evaporites (Cooper et al., 1995;
Sarmiento-Rojas et al., 2006). There are few Mesozoic-Cenozoic
plutons intruding the Eastern Cordillera, but the Santander Massif
of the northernmost Eastern Cordillera contains Jurassic quartz

249

monzonite, diorite, and granite of calc-alkaline and potassic afnity


(Aspden et al., 1987).
2.2. Nuevo Mundo Syncline
West-vergent thrust faults and anticlineesyncline pairs dene
the transition between the Magdalena Valley and Eastern
Cordillera (Butler and Schamel, 1988; Colletta, 1990; Dengo and
Covey, 1993; Restrepo-Pace et al., 2004). The Nuevo Mundo
Syncline exposes all Cenozoic units of the northern Middle Magdalena Valley Basin. This nearly symmetric, north-plunging
syncline is bounded by the west-vergent La Salina Thrust Fault
to the west and the Los Cobardes Anticline to the east (Fig. 2). This
study focuses on exposures near the type sections of the Paleogene Lisama, La Paz, and Esmeraldas Formations (Morales et al.,
1958; Schamel, 1991; Gmez et al., 2005b). Although the Middle
Magdalena Valley unconformity of mainly Eocene age is well
documented farther west (Gmez et al., 2003, 2005b), available
lithostratigraphic and chronostratigraphic constraints for the
Nuevo Mundo Syncline indicate relatively continuous Cenozoic
accumulation along the eastern limb (Pardo-Trujillo et al., 2003;
Pardo-Trujillo, 2004) with discontinuous accumulation along the
western limb. Importantly, map relationships and seismic reection data (Figs. 2 and 3) show that the Eocene La Paz Formation
along the eastern limb of the syncline thins systematically westward in the subsurface and may be absent farther west in the axis
of the Middle Magdalena Valley Basin (Gmez et al., 2005b;
Caballero, 2010; Caballero et al., 2010).
The Paleocene Lisama Formation marks the main transition
from marine (Maastrichtian Umir Formation) to nonmarine
(Eocene La Paz Formation) sedimentation, with thicknesses of
820e1225 m (Ramirez, 1988). Surface exposures show the thickness of the Lisama Formation to be nearly uniform across the Nuevo
Mundo Syncline. The Lisama, which conformably overlies the Umir
Formation, is dened as Paleocene age by palynomorph fossil
assemblages (Pardo-Trujillo and Jaramillo, 2002; Pardo-Trujillo
et al., 2003; Pardo-Trujillo, 2004).
An earlyemiddle Eocene age for the base of the La Paz Formation is dened on the basis of palynological age constraints for the
eastern limb of the syncline (Pardo-Trujillo and Jaramillo, 2002;
Pardo-Trujillo et al., 2003; Pardo-Trujillo, 2004). The La Paz
Formation ranges in thickness from 1000 to 1280 m in the eastern
limb to 0e90 m near the western limb (Ramirez, 1988; Caballero,
2010). This westward thinning, and local pinchout, of the La Paz
Formation across the syncline is also expressed in seismic reection
data (Figs. 2 and 3). At the surface, the La Paz is not present west of
the La Salina Thrust and its subsurface extent is questionable. The
La Paz has a lithostratigraphic equivalent, the Cantagallo Sandstone,
along the western ank of the Middle Magdalena Valley in the
footwall of the east-vergent Cantagallo Thrust Fault (Fig. 2B), but
age correlations are lacking (Surez, 1997; Gmez et al., 2005b). In
the eastern limb of the syncline, the La Paz Formation consists of
nonmarine facies and has a lower ne-grained and upper coarsegrained unit discussed in more detail below.
The upper EoceneeOligocene Esmeraldas Formation contains
mixed mudstone and sandstone facies similar to those of the
lower La Paz Formation and is about 1200 m thick at its type
locality (Morales et al., 1958). Although the Esmeraldas Formation
is present across the basin, it thins westward across the Nuevo
Mundo Syncline (Figs. 2 and 3). The Esmeraldas conformably
overlies the La Paz Formation in most locations, but sits directly
on the lower La Paz and/or Lisama along the western margin of
the syncline, in close proximity to the La Salina Thrust (Caballero,
2010; Caballero et al., 2010). Age control from fossil palynomorph
assemblages and bivalves and gastropods of the Los Corros fossil

250

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

Fig. 3. Seismic reection prole and surface topography across the Nuevo Mundo Syncline (see Figs. 1 and 2 for location) showing interpreted Cenozoic stratigraphic units and
approximate depths. Note the signicant westward thinning of the La Paz Formation and moderate westward thinning of the Esmeraldas Formation.

horizon (which caps the Esmeraldas) dene a depositional age of


late Eocene to early Oligocene (Ramirez, 1988; Gmez et al.,
2005b).
3. Sedimentology
Sedimentological descriptions and interpretations (Tables 1 and
2) are based on 11 measured stratigraphic sections in the eastern
and western limbs of the Nuevo Mundo Syncline. The thickest
section (Fig. 4) was measured along the Sogamoso River (Bucaramanga-Barrancabermeja highway) in the eastern limb of the
syncline (Fig. 2A). Other workers have measured and published
lower-resolution stratigraphic columns from similar locations
(Pardo-Trujillo and Jaramillo, 2002; Pardo-Trujillo et al., 2003;
Pardo-Trujillo, 2004; Gmez et al., 2005b). Limited lateral and
vertical exposures in places locally hinder assessments of architectural elements and stratigraphic relationships among lithofacies
associations.
Our descriptions focused on the eld characteristics and
occurrences of different lithofacies, building upon previously
established lithofacies codes (e.g., Miall, 1977, 1985, 1996; Smoot,
1991; Uba et al., 2005). Table 1 identies 15 different lithofacies,
including 1 conglomerate (Gc), 9 sandstone (Sml, Smt, St, Sh, Sl, Sr,
Srs, Srf, Srw), 4 mudstone (Fl, Frw, Fm, Fps), and 1 coal (C) lithofacies. These lithofacies (Figs. 5 and 6) are arranged into various
packages and categorized into 5 different lithofacies associations
(Table 2).
3.1. Lithofacies association 1: upward coarsening ripple-laminated
sandstone and interbedded mudstone
Lithofacies association 1 is dened by up to 25 m intervals of
upward coarsening and thickening packages of symmetric ripplelaminated and massive tabular sandstone with interbedded
mudstone (Fig. 5A and B). This association is limited to the lower
Lisama Formation, and well represented in the lowermost 50 m of
the measured section (Fig. 4). The brown to yellow, moderately to
well sorted, very ne- to medium-grained sandstone beds thicken
upsection from 210 mm at the base to 1e1.7 m at the top.

Mudstones are planar laminated, range from clay to silt size, and
commonly show ute casts at their bases. Occasionally, strata of
lithofacies association 1 are scoured by shallow, cross-stratied
sandstones with lenticular geometries. Ramirez (1988) observed
arenaceous foraminifera and coal seams within beds of this lithofacies association.
We attribute deposition of lithofacies association 1 to the delta
slope to delta front of a wave-dominated delta in a marginal marine
environment. Marine inuence is deemed signicant by the
presence of arenaceous foraminifera (Linke and Lutze, 1993) and
symmetric ripples suggestive of wave inuence. These ripples and
the lack of hummocky cross-stratication indicate wave-inuenced
deposition under relatively weak oscillatory ows (Tye et al., 1999;
Willis and Gabel, 2001). Thin (cm-scale) tabular sandstone beds
formed as a result of wave reworking and discontinuous storm
deposition. Mudstones represent periods of suspension fallout
during reduced ow conditions, which commonly resulted in mud
draping of abandoned sands. The upward coarsening and thickening pattern, and incision by lenticular trough cross-stratied
sandstones, indicate progradation of a delta slope to delta front
with distributary channels feeding subaqueous mouth bars (Tye
et al., 1999; Bhattacharya and Giosan, 2003).
3.2. Lithofacies association 2: upward-thickening ripple, wavy, and
cross-stratied sandstone
Gray to brown, very ne- to ne-grained sandstone beds
dene lithofacies association 2 (Fig. 5C and D), which is best
expressed in the Lisama Formation (notably at the 340e370 m
level of the measured section; Fig. 4), commonly overlying
deposits of lithofacies association 1. Bed thicknesses range from
0.02 to 2.5 m, and generally thicken upsection within intervals up
to 25 m thick. Most beds are tabular and contain current or wave
ripple lamination and aser or lenticular bedding. Trough crossstratied sandstone beds exhibit erosive bases and broad lenticular geometries.
We interpret strata of lithofacies association 2 to have been
deposited principally in a subaerial delta plain. Thin to medium
beds of ripple and trough cross-stratied sandstone are consistent

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

251

Table 1
Description and interpretation of observed sedimentary lithofacies (after Miall, 1985; Uba et al., 2005).
Facies code

Description

Interpretation

Gc

Massive, poorly sorted, clast-supported, imbricated, polymictic, subrounded granule to cobble


conglomerate; interval thickness range: 0.5e10 m
Massive, lens-shaped, poorly to well sorted, subrounded to subangular, ne- to very coarse-grained
sandstone; interval thickness range: 0.4e4 m
Massive, tabular bedded, poorly to well sorted, subrounded to subangular, very ne- to very coarsegrained sandstone; interval thickness range: 0.01e7 m

Traction transport of gravel


bedload in uvial channels
Rapid sand deposition in uvial
or mouth bar channels
Rapid sheetow deposition of sand
with limited channelization,
possible bioturbation
Migration of 3D sand dunes in
uvial or mouth bar channels
Upper ow regime plane-bed deposition
of sand in channels or crevasse splays
Lower ow regime plane-bed
deposition of sand
Sandy ripple migration in
channels or crevasse splays
Sandy 2D wave ripples in oscillatory ows

Sml
Smt

St
Sh
Sl
Sr
Srs
Srf

Trough cross-stratied, poorly to well sorted, subrounded to subangular, ne- to coarse-grained


sandstone; interval thickness range: 0.3e11 m
Horizontally stratied, moderately to well sorted, subrounded to subangular, ne- to mediumgrained sandstone; interval thickness range: 0.1e2 m
Planar laminated, moderately to well sorted, subangular, very ne- to medium-grained sandstone;
interval thickness range: 0.15e4 m
Ripple cross-stratied, occasional climbing ripples, moderately sorted, very ne- to ne-grained
sandstone; interval thickness range: 0.3e3 m
Symmetric ripple-laminated, well sorted, very ne- to ne-grained sandstone; interval thickness
range: 0.1e2 m
Flaser or lenticular bedded, occasional climbing ripples, moderately sorted, ne-grained sandstone;
interval thickness range: 0.2e0.5 m

Fl

Wavy laminated, well sorted, very ne- to ne-grained sandstone; interval thickness range: 0.3
e6 m
Planar laminated claystone to siltstone; interval thickness range: 0.01e3.2 m

Fm

Massive claystone to siltstone; interval thickness range: 0.01e36 m

Frw

Wavy laminated siltstone; interval thickness range: 0.2e2.5 m

Fps

Massive, moderately developed paleosols and pedogenic nodules in siltstone; interval thickness
range: 0.5e15 m
Coal, plant remains, carbonaceous mudstone

Srw

with deposition in proximal segments of shallow distributary


channels (Galloway, 1976; Orton and Reading, 1993). Very negrained sandstones with lenticular or aser bedding indicate
deposition in the distal segments of distributary channels with tidal

Sandy ripple migration and suspension


settling of mud during alternating
ow conditions
Rapid deposition of sandy ripples
Suspension fallout of mud in delta slope
or uvial overbank setting
Suspension fallout of mud in delta slope
or uvial overbank setting
Muddy ripple migration in delta slope or
uvial overbank setting
Soil development in abandoned channel
or overbank setting
Poorly developed overbank accumulation
of organic matter

inuence. The latter facies of the lower delta plain were deposited
in close proximity to delta front deposits of lithofacies association 1,
accounting for the transitional nature between associations 1 and 2
within the Lisama Formation.

Table 2
Lithofacies associations and interpretations.
Facies association

Lithofacies

Description

Stratigraphic occurrence

Interpretation

1: Upward coarsening
ripple-laminated
sandstone and
interbedded mudstone

Sr, Smt,
Fm, Fl, St

Lower Lisama Formation

Delta front to delta slope of


a wave-dominated delta,
including distributary
mouth bars

2: Upward-thickening
ripple, wavy and crossstratied sandstone

Sr, Srw, Sl,


Srf, St, Srs

Lisama Formation

Subaerial delta plain,


including distributary
channels

3: Thick bedded, trough


cross-stratied and
massive sandstone

St, Sml,
Smt, Sh

Throughout La Paz Formation


(amalgamated in upper 230 m);
limited occurrence in Esmeraldas
Formation

Principally braided uvial


channel, including possible
uvial megafan (upper
230 m of La Paz Formation)

4: Cross-stratied
sandstone with local
conglomerate

St, Smt, Sl,


Gc, Fm

Lowermost La Paz Formation


(conglomerate in lower 100 m);
throughout Esmeraldas Formation

Meandering uvial channel


deposition

5: Thin bedded, massive


to laminated mudstone
with interbedded
sandstone

Fm, Frw, Fl
Sh, Sr, Fps, C

Up to 25 m intervals of brown to yellow,


symmetric ripple-laminated, very ne- to
medium-grained sandstone with brown,
planar laminated claystone to siltstone;
sandstone bed thicknesses increase from
2 to 10 mm at base to 1e1.7 m at top of
interval; occasional lenticular crossstratied sandstone beds
Gray to brown, tabular beds of aser/
lenticular bedded, wave/current rippled,
or trough cross-stratied very ne- to negrained sandstone; 0.02e2.5 m beds
commonly thicken upsection
White to yellow, ne- to very coarse-grained
sandstone; 0.5e10 m intervals of wedge and
lens-shaped beds with erosional basal
surfaces; beds occasionally contain pebbles
and mud rip-up clasts near base
Up to 0.5e3 m beds of brown to yellow,
ne- to very coarse-grained sandstone;
local 0.5e5 m intervals of normally
graded conglomerate
0.5e30 m intervals of gray-purple to black,
massive to wavy laminated, normally graded
claystone to very ne-grained sandstone;
interbeds of horizontal and ripple stratied,
normally graded sandstone with mudstone
rip-up clasts

Lower La Paz Formation;


throughout Esmeraldas
Formation

Fluvial oodplain with


deposition of overbank
mud and crevasse splay
sand

252

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

Fig. 4. Composite measured stratigraphic section of the Paleogene succession exposed along the eastern limb of the Nuevo Mundo Syncline (see Fig. 2 for location) showing
lithofacies types, sedimentary structures, paleocurrent orientations, sample locations, and eld measurements of gamma ray response (CPS counts per second). Paleocurrent
indicators with single arrowheads represent stations within measured sections; double arrowheads indicate data projected from nearby stations.

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

253

Fig. 5. Photographs representing lithofacies in the Lisama Formation. (A) Upward coarsening, interbedded sandstone (Srs, Smt) and mudstone (Fm) of lithofacies association 1
(delta front to delta slope). Beds coarsen and thicken upsection to the right. White box indicates location of (B), which shows interbedded sandstone (Srs, Smt) and siltstone (Fm, Fl)
in detail. Rock hammer (lower right) for scale. (C) Upward-thickening, ripple (Sr, Srf), wavy (Srw), and cross-stratied (St) sandstone of lithofacies association 2 (subaerial delta
plain). Person (lower left) for scale. White box indicates location of (D), which shows ripple (Sr) and aser/lenticular bedded (Srf) sandstone. Pencil (lower left) for scale.

3.3. Lithofacies association 3: thick bedded, trough cross-stratied


and massive sandstone
Lithofacies association 3 (Fig. 6A and B) represents the principal
occurrences of thick bedded, trough cross-stratied and massive
sandstones throughout the La Paz Formation and in selected
intervals of the Esmeraldas Formation. This lithofacies association
becomes amalgamated in the upper 230 m of the La Paz Formation
(Fig. 4). Sandstones are white to yellow with grain sizes ranging
from ne to very coarse. Trough cross-stratication is the dominant
sedimentary structure. Beds range from 0.5 to 10 m thickness and
are commonly wedge or lens-shaped with sharply erosional basal
surfaces, although several-m-thick intervals of tabular beds are
present locally. The bases of many beds contain pebbles and
mudstone rip-up clasts. Upward ning trends are common in
individual beds and in stacked packages of massive to crossstratied sandstone up to 30 m thick.
Lithofacies association 3 is interpreted as individual and stacked
channel-ll deposits of a braided uvial system. Intervals containing amalgamated sandstones with large lens-shaped macroforms
devoid of mudstone and crevasse-splay deposits are interpreted to
represent relatively deep, perennial, sandy braided streams. Other
supporting characteristics of braided streams include pebble lags
and scour surfaces with mudstone rip-up clasts (Miall, 1977).
Upward ning trends suggest either waning energy near the end of
individual depositional events or diminished ow related to
systematic avulsion or migration of the main channel away from
the locality. Local stacks of tabular beds within the same succession
also suggest the presence of shallower, braided streams (Miall,
1996; Uba et al., 2005), potentially as broad sheetow complexes
(Hampton and Horton, 2007). We speculate that some

amalgamated sandstone intervals may represent rapid channel and


sheetow deposition in the medial portions of uvial megafans
(DeCelles and Cavazza, 1999; Horton and DeCelles, 2001; Uba et al.,
2005; Hampton and Horton, 2007).
3.4. Lithofacies association 4: cross-stratied sandstone with local
conglomerate
Lithofacies association 4 contains 0.5e3 m intervals of crossstratied sandstone and subordinate 0.5e5 m intervals of normally graded conglomerate (Fig. 6C). The conglomerates are limited
to the lower 100 m of the La Paz Formation (Fig. 4) and consist of
clast-supported, imbricated, polymictic conglomerates with subrounded pebbles and cobbles (Fig. 6C). Where sufciently exposed,
the conglomerates have erosive basal surfaces and broadly lenticular geometries. The sandstones of this lithofacies association are
common throughout the Esmeraldas Formation and include brown
to yellow, trough cross-stratied and massive beds up to 0.5e3 m
thick. These sandstones exhibit tabular and lateral accretion surfaces
(Fig. 6D) and are commonly separated by 0.5e30 m thick intervals of
massive to laminated siltstone of lithofacies association 5.
Cross-stratied sandstone and local normally graded conglomerate of lithofacies association 4 are considered the product of
traction transport of sand and gravel bedload within meandering
uvial channels. The trough cross-stratied sandstones display
lateral accretion surfaces and scour into or are capped by thin
bedded, laminated to massive mudstone and interbedded sandstone
of lithofacies association 5 (Fig. 6D), suggesting a meandering stream
environment (Allen, 1965; Miall, 1977). Conglomerate clast imbrication suggests lateral or longitudinal bars and upward ning indicates deposition under waning energy. The erosive basal scours for

254

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

Fig. 6. Photographs representing lithofacies in the La Paz and Esmeraldas Formations. (A) Amalgamated, lens-shaped (Sml), tabular (Smt), and cross-stratied (St) sandstone beds of
lithofacies association 3 (braided uvial channel) in the upper La Paz Formation. Arrows indicate the base of a single lens-shaped channel. Truck (lower right) for scale. (B) Trough
cross-stratied sandstone (St) of lithofacies association 3 (braided uvial channel) in the lower La Paz Formation. Two persons (upper right) for scale. (C) Interbedded crossstratied sandstone (St) and conglomerate (Gc) of lithofacies 4 (meandering uvial channel). Arrow indicates contact. Rock hammer (center) for scale. (D) Interbedded lithofacies associations 4 and 5 (meandering uvial channel and uvial oodplain): tabular and lens-shaped beds of massive (Sml, Smt) and trough cross-stratied (St) sandstone
overlying massive (Fm) and laminated (Fl) mudstone. Two persons (center, lower right) for scale. (E) Overbank laminated mudstone (Fl) with interbedded sandstones (Sr, Sh) of
lithofacies 5 (uvial oodplain). Person (upper left) for scale.

imbricated conglomerate beds, presence of interbedded trough


cross-stratied sandstones with lateral accretion surfaces, and
occurrence with lithofacies association 5 are consistent with a mixed
gravel and sand meandering uvial system (Miall, 1996; Nanson and
Knighton, 1996).
3.5. Lithofacies association 5: massive to laminated mudstone with
interbedded sandstone
Lithofacies association 5 consists of gray-purple to black
mudstones concentrated in the lower 700 m of the La Paz Formation and throughout the Esmeraldas Formation (Fig. 4), commonly
interbedded with lithofacies association 4. The thin bedded

(mm- to cm-scale) mudstones occur in 0.5e30 m thick intervals,


show massive or wavy laminated texture, and exhibit normal
grading from clay to very ne-grained sand (Fig. 6D). Interbedded
with some mudstones are laterally extensive sheets of horizontal
and ripple stratied, normally graded sandstones containing
mudstone rip-up clasts (Fig. 6E). These interbedded sandstones are
uniformly thin- to medium-bedded (5e40 cm thick) and tabular,
with no examples of thick, stacked lenticular sandstones. Thin
(1e20 cm) beds of coal are present locally.
We interpret the mudstone-dominated deposits of lithofacies
association 5 to represent overbank sedimentation in a uvial
oodplain setting (Miall, 1977, 1996; Gmez et al., 2005b; Uba et al.,
2005). These overbank mudstones are interbedded with thin,

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

laterally extensive sandstone sheets with ripple and horizontal


stratication interpreted as crevasse-splay deposits (Miall, 1996;
Horton and DeCelles, 2001). The thick mudstone intervals of lithofacies association 5 separating sandstone channels of lithofacies
association 4 are suggestive of high suspended sediment loads
(Smoot, 1991). Moreover, the close association of lithofacies associations 4 and 5 and lack of protracted sand deposition in channel
settings may suggest overbank deposition in meandering rather
than anastomosing uvial systems (e.g., Makaske, 2001).
4. Sediment provenance
4.1. Sediment dispersal patterns
Paleocurrent indicators were measured along both the eastern
and western limbs of the Nuevo Mundo Syncline (Fig. 4), with
compiled paleocurrent data categorized by formation (Fig. 7).
Paleocurrents were determined from trough cross limbs (method I
of DeCelles et al., 1983), conglomerate clast imbrications, and ute
cast orientations. All data were corrected for the 10e20 northward
fold plunge and 20e60 dip of strata on the limbs of the syncline.
A total of 56 ute casts near the base of the Lisama Formation
show a dominant northward paleoow (Fig. 7A). Upsection, Lisama
paleocurrents switch to dominantly east-directed (Fig. 7B). This
change in sediment dispersal is consistent with an observed shift in
UePb age spectra from Proterozoic to Phanerozoic ages in corresponding levels of the section (Nie et al., 2010) and an increase in
volcanic lithic fragment content (discussed below). The shift

255

potentially indicates a change in the dominant sediment source


region from the Guyana Shield (Amazonian Craton) to the Central
Cordillera.
For the upper Lisama Formation and La Paz Formation, trough
cross-stratication and clast imbrication measurements show
principally east-directed paleocurrents. For the lower La Paz section
(Fig. 7C), these results contrast with those of Gmez et al. (2005b)
who recorded west-directed paleocurrents. Although the present
study collected substantially more measurements in the lower La
Paz, some degree of scatter may be the product of spatial variability
in ow orientations measured in different locations.
Although an apparent change in source proximity is observed
from the ne-grained lower La Paz to coarse-grained upper La Paz,
a consistent eastward paleoow pattern persists throughout the La
Paz Formation. Whereas the lower La Paz is dominated by
meandering uvial deposits characteristic of lithofacies associations 4 and 5, the upper La Paz Formation is dominated by moreproximal, coarser grained braided uvial deposits of lithofacies
association 3. The consistent paleocurrents through this pronounced
facies shift suggests the growing inuence of the La Cira-Infantas
paleohighs during accumulation of upper La Paz sediments. These
western paleohighs potentially shut off transverse rivers sourced
from the more-distal Central Cordillera and directly sourced new,
shorter rivers more-proximal to the Nuevo Mundo Syncline.
Near the base of the Esmeraldas Formation, in the eastern limb
of the Nuevo Mundo Syncline, paleocurrents from trough crossstrata show a switch from east-directed to dominantly westdirected ow (Fig. 7D). This reversal, also observed in the western
limb (Fig. 7E), is nearly coeval with a shift toward more-distal
meandering rivers represented by the combination of lithofacies
associations 4 and 5. Contemporaneous with the aforementioned
changes, detrital zircon UePb age spectra show the elimination of
Jurassic-Early Cretaceous (150e100 Ma) ages associated with
a Central Cordilleran source and an increase in Grenville
(1200e900 Ma) ages associated with an Eastern Cordilleran source.
Collectively, these different datasets reveal a latest Eocene change
in Nuevo Mundo provenance from western sources such as the
Central Cordillera and La Cira-Infantas paleohighs to eastern sources such as the axial to western margin of the Eastern Cordillera.
4.2. Sandstone compositions
To further assess Paleogene source areas for the Middle Magdalena Valley Basin, medium-grained sandstone samples for
Table 3.
Parameters for sandstone petrographic point counts.

Fig. 7. Rose diagrams displaying paleocurrent data, including the vector mean and
standard deviation, vector magnitude, maximum percentage, number of measurements, and number of stations. Unless otherwise noted, all measurements are from the
eastern limb of the Nuevo Mundo Syncline. (A) Flute cast measurements from the base
of Lisama Formation. (B) Trough cross-stratication measurements from the upper
Lisama Formation. (C) Measurements of trough cross-stratication and clast imbrication in the La Paz Formation. (D) Trough cross-stratication measurements from the
Esmeraldas Formation. (E) Trough cross-stratication measurements from the
Esmeraldas Formation along the western limb of the syncline.

Symbol

Grain Categories

Calculated Parameters

Qm
Qp

monocrystalline quartz
polycrystalline quartz

Qpt
Fp
Fk
Lms
Lmp
Lmsc
Lvc

polycrystalline quartz with tectonic fabric


plagioclase feldspar (including albite)
potassium feldspar
slate lithic fragments
phyllite lithic fragments
schist lithic fragments
volcaniclastic lithic fragments

Q-F-L:
Q Qm Qp Qpt
Lch
F Fp Fk
L Lm Lv Ls

Lvl
Lvf
Lch

lathwork volcanic lithic fragments


felsitic volcanic lithic fragments
chert lithic fragments

Lss
Lsc
M
D

siltstone lithic fragments


claystone lithic fragments
monocrystalline mica
heavy minerals

Qm-F-Lt:
Q Qm
F Fp Fk
L Lm Lv Ls
Lch Qp Qpt
Lm-Lv-Ls:
Lm Lms Lmp
Lmsc
Lv Lvc Lvl Lvf
Ls Lch Lss Lsc

256

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

petrographic analysis were collected in the Nuevo Mundo Syncline.


Each sample was cut into a standard thin section and stained for
plagioclase and potassium feldspar, but not injected with blue dye.
A total of 42 thin sections were analyzed from the measured section
along the eastern limb (Fig. 4) and 13 thin sections from the
western limb. Samples were counted according to the GazziDickinson method (e.g., Gazzi, 1966; Dickinson, 1970; Ingersoll
et al., 1984), with 450 points counted per sample. All points
counted represent framework sand grains (>0.0625 mm) in which
each counted grain is placed into one of the 16 compositional
categories (3 quartz, 2 feldspar, 9 lithic fragments, 2 other minerals)
listed in Table 3. Representative grains are shown in Fig. 8. During
the counting process, if the microscope crosshairs landed on
matrix, the point was not counted. If the crosshairs landed on
authigenic mineral growth, calcite cement, or quartz cement, the
original underlying grain was counted. This method of point

counting does not account for potential loss of feldspar through


diagenesis; therefore we focus our interpretations on quartz and
lithic distributions (Milliken, 1988, 1992; Milliken et al., 1989).
Point-count results for each sample (Table 4) are recalculated to
assess normalized percentages of quartz-feldspar-lithic fragments
(Q-F-L %), monocrystalline quartz-feldspar-total lithic fragments
(Qm-F-Lt %) and metamorphic-volcanic-sedimentary lithic fragments (Lm-Lv-Ls %). The upsection stratigraphic trends in sandstone compositions are most clearly represented in Qm-F-Lt and
Lm-Lv-Ls ternary diagrams for both limbs of the Nuevo Mundo
Syncline (Fig. 9). Nearly all samples contain strained monocrystalline quartz (Qm), polycrystalline quartz (Qp), and polycrystalline quartz with tectonic fabric (Qpt), none of which exhibit
obvious upsection trends in occurrence. In addition, grain shapes in
nearly all samples fall in the range of subangular to subrounded
with no obvious variations among formations.

Fig. 8. Photomicrographs of sandstone petrographic thin sections from the eastern limb of the Nuevo Mundo Syncline. (A) RS014P from upper Lisama Formation. (B) RS015P from
upper Lisama Formation. (C) RS475P from lower La Paz Formation. (D) RS481P from lower La Paz Formation. (E) SOG081010 from upper La Paz Formation. (F) SOG08113 from
Esmeraldas Formation.

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

257

Table 4
Modal sandstone point-count data from the eastern and western limbs of the Nuevo Mundo syncline. Italics represent stratigraphic levels projected from other nearby sections.
Sample

Level(m)

Formation

Eastern limb, Nuevo Mundo syncline (n 42)


1SOG14
33
Lisama
RS014P
1125
RS015P
1135
RS016P
1138
RS020P
1232
lower La Paz
RS021P
1238
RS022P
1250
U08022
1260
RS023P
1274
1SOG21
1300
RS066P
1345
RS095P
1422
RS096P
1443
RS102P
1527
RS104P
1547
RS148P
1642
RS472P
1733
RS475P
1764
RS476P
1786
RS481P
1786
RS484P
1809
RS599P
1839
CU603P
1981
upper La Paz
RS604P
1995
CU607P
2001
CU610P
2015
CU612P
2039
SOG08102
2072
SOG08103
2091
RS613P
2095
SOG08105
2103
SOG08106
2112
SOG08108
2141
RS614P
2149
SOG081010
2162
RS625P
2312
Esmeraldas
SOG08113
2520
RS758P
2543
SR0106091
2600
RS487P
2797
RS530P
2818
RS598P
3275
Western limb, Nuevo Mundo syncline (n 13)
WS0107091
base
Lisama
PUT08141
La Paz
PUT081410
PUT08132
WS0107094
Esmeraldas
PUT08148
PUT08147
PUT08146
PUT08145
WS0110097
PUT08143
PUT08142
WS0109095
top

Number

Qm

Lt %

Qm

Lt %

Lm-Lv-Ls %

Qm

Lt

Lm

Lv

Ls

1
2
3
4
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
1
2
3
4
5
6
7
8
9
10
11
12
13
1
2
3
4
5
6
7

69
56
59
53
65
62
70
87
75
74
81
69
57
48
70
61
82
76
67
64
76
75
90
92
97
91
87
86
90
90
77
80
89
91
89
87
94
86
83
83
89
92

6
5
2
4
2
4
3
0
3
5
1
3
7
11
6
9
1
3
8
5
12
9
1
0
0
0
0
5
1
1
5
4
1
5
0
6
2
3
5
5
5
3

25
39
39
43
33
34
27
13
22
21
18
28
36
41
24
30
17
21
25
31
12
16
9
8
3
9
13
9
9
9
18
16
10
4
11
7
4
11
13
12
6
5

54
47
52
49
59
54
64
59
71
60
74
66
50
42
65
58
70
67
62
56
70
68
83
81
95
83
81
78
82
83
62
63
76
86
82
84
81
76
68
79
75
86

6
5
2
4
2
4
3
0
3
5
1
3
7
11
6
9
1
3
8
5
12
9
1
0
0
0
0
5
1
1
5
4
1
5
0
6
2
3
5
5
5
3

40
48
46
47
39
42
33
41
26
35
25
31
43
47
29
33
29
30
30
39
18
23
16
19
5
17
19
17
17
16
33
33
23
9
18
10
17
21
28
16
20
11

41
54
50
39
45
33
29
57
23
47
29
38
47
20
26
41
48
37
27
36
34
45
51
51
39
45
34
40
37
44
67
68
56
58
41
34
63
56
73
65
69
41

25
38
43
52
44
48
57
5
57
16
46
46
39
63
58
49
38
43
60
40
49
40
19
16
28
19
14
13
16
16
6
7
7
15
7
17
4
9
2
5
4
19

34
8
7
9
11
19
14
38
20
37
25
16
14
17
16
10
14
20
13
24
17
15
30
33
33
36
52
47
47
40
27
25
37
27
53
49
33
35
25
11
27
41

1
1
2
3
1
2
3
4
5
6
7
8
9

78
98
95
97
82
78
79
80
73
72
72
77
69

0
0
1
1
6
5
8
6
6
7
9
6
18

22
2
4
2
12
17
13
14
21
20
19
17
13

53
95
85
88
66
71
65
70
60
56
62
67
48

0
0
1
1
6
5
8
6
6
7
9
6
18

47
5
14
11
28
24
27
24
34
36
29
27
34

58
62
70
73
58
57
64
48
54
52
61
60
48

9
4
5
2
9
10
4
3
5
12
10
7
17

33
34
25
25
33
33
32
49
41
36
29
33
35

The compositional data presented here for 55 thin sections


(5 Lisama, 34 La Paz, 16 Esmeraldas) broadly agree with the
compositional data presented by Gmez et al. (2005b) for
9 samples from corresponding levels (2 Lisama, 5 La Paz, 2 Esmeraldas). Our data, however, indicate greater degrees of compositional variability and upsection compositional shifts from mainly
litharenites and minor feldspathic litharenites to sublitharenites
and subordinate quartzarenites (Fig. 9).
4.2.1. Lisama Formation
Four Lisama litharenites (sandstone classication of Folk, 1980)
were point counted from the eastern limb of the syncline (Fig. 9A

and B) and one sublitharenite from the western limb. Eastern limb
samples exhibit a mean composition of Qm51-F4-Lt45. Lisama feldspars are 94% plagioclase (pink-stained Ca-rich feldspar) with
common albite twinning. Lithic fragments show a distinction
between a basal Lisama sample (Lm41-Lv25-Ls34) and three samples
near the top (Lm48-Lv44-Ls8). Sedimentary lithic fragments (Ls) are
less prevalent in the uppermost Lisama, where more volcanic
felsitic (Lvf), volcanic lathwork (Lvl), volcaniclastic (Lvc), and
metamorphic (Lm) lithic fragments are present. The uppermost
samples also show a greater proportion of higher grade metamorphic fragments such as schist (Lmsc). All samples show
a proportion of polycrystalline quartz (Qp) and polycrystalline

258

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

Fig. 9. Qm-F-Lt and Lm-Lv-Ls sandstone ternary diagrams for 12 samples from the western limb (left) and 42 samples from the eastern limb (right) of the Nuevo Mundo Syncline.
(A, C) Gray arrows show upsection increases in total lithic (Lt) and sedimentary lithic (Ls) content for the western limb. (B, D) For the eastern limb, gray arrows show an upsection
decrease in total lithic (Lt) content and a corresponding increase then decrease in the relative proportion of volcanic lithic (Lv) to sedimentary lithic (Ls) content.

quartz with tectonic fabric (Qpt). The composition of a single


Lisama sample from the western limb is also markedly different,
containing no feldspar and signicantly less volcanic lithic fragments (Lm58-Lv9-Ls33). However, the limited thickness of the
mapped Lisama in the western limb (Fig. 2A) (Caballero, 2010;
Caballero et al., 2010) precludes accurate assignment of this
sample (WS0107091) within the Lisama Formation.
4.2.1.1. Interpretations. Although limited, the ve Lisama samples
reveal an upsection shift toward dominantly volcanic and metamorphic lithic fragments. This trend and accompanying reduction
in sedimentary lithic fragments is consistent with the introduction
of magmatic-arc and metamorphic source areas, as expected for
contributions from the intrusive and extrusive igneous rocks and
metamorphic host rocks of the Central Cordillera (Aspden and
McCourt, 1986). The San Lucas Range along the northwestern
margin of the Middle Magdalena Valley Basin (Fig. 1) is a further
possible source for volcaniclastic grains ultimately derived from
magmatic-arc rocks of principally Mesozoic age from the Central
Cordillera (Kammer and Snchez, 2006). The lack of feldspar
content, which is a key component of many magmatic-arc sources
(Marsaglia and Ingersoll, 1992), could be due to selective dissolution of feldspar grains during diagenesis (Milliken et al., 1989).
Schist fragments may be linked to a moderate- to high-grade belt of
Paleozoic metamorphic rocks in the Central Cordillera (McCourt
et al., 1984; Aspden and McCourt, 1986). This compositional
evidence supports the paleocurrent data (reported above) and
detrital zircon UePb ages (Nie et al., 2010; Saylor et al., 2011)
suggesting a mid-Paleocene onset for widespread sedimentation
derived from early shortening-induced exhumation of igneous and
metamorphic rocks of the Central Cordillera.
4.2.2. Lower La Paz Formation
Based on an upsection facies shift to mostly cliff-forming
sandstone (Fig. 4), a stratigraphic boundary may be assigned
between the lower La Paz (lower 600e700 m) and upper La Paz
(upper 300e400 m). For the lower La Paz, 18 samples were point
counted, yielding mostly litharenite with subordinate sublitharenite and feldspathic litharenite compositions and a mean
composition of Qm62-F5-Lt33 (Fig. 9C and D). Overall, 63% of

feldspars in this group of samples are plagioclase. Although the


proportions remain low, the upper 9 samples show a doubling of
mean feldspar content relative to the lower 9 samples. The
proportion of lithic fragments decreases upsection through the
lower La Paz with a mean lithic composition of Lm37-Lv44-Ls19.
4.2.2.1. Interpretations. The principally litharenitic sandstone
compositions for the lower La Paz Formation are generally
comparable to the upper Lisama Formation. Volcanic and metamorphic compositions remain the dominant lithic fragments,
consistent with a continued western provenance in which the
Central Cordillera remained the dominant source for sediment
deposited in the Middle Magdalena Valley Basin. A modestly
reduced proportion of lithic fragments and upsection increase in
feldspar content, although consistently <15%, could reect relatively greater input from igneous sources.
4.2.3. Upper La Paz Formation
The 13 sandstone samples of the upper La Paz Formation from
the eastern limb of the syncline record a shift to higher mineralogical maturity (sublitharenite and minor quartzarenite) with
a mean composition of Qm80-F2-Lt18 (Fig. 8E). There is a signicant
decrease in both feldspar and lithic fragments at the boundary
between the lower and upper La Paz (Fig. 9; Table 4) and nearly all
feldspar grains in the upper La Paz are plagioclase. Relative to lower
levels of the section, the proportions of volcanic lithic fragments
decrease and sedimentary lithic fragments increase. The total lithic
compositions, with a mean value of Lm49-Lv14-Ls37, are considerably enriched in sedimentary lithic fragments at the expense of
volcanic lithic fragments.
Along the western limb, 3 samples were collected from stratigraphic levels mapped as La Paz Formation. The total stratigraphic
thickness in this portion of the La Paz Formation is substantially less
than any part of the eastern limb (Fig. 2) (Caballero, 2010; Caballero
et al., 2010). These samples are quartzarenites with mean compositions of Qm96-F1-Lt3 and Lm68-Lv4-Ls28.
4.2.3.1. Interpretations. Variations in the proportion and composition of lithic fragments and feldspar suggest a different source area
for upper Eocene deposits of the upper La Paz Formation. Lower

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

proportions of volcanic rock fragments indicate a diminished


inuence of the Central Cordillera on detrital compositions. Indeed,
the enhanced maturity reected by the higher quartz content and
higher ratio of sedimentary to volcanic lithic fragments may
suggest erosional recycling of a mature Cretaceous stratigraphic
succession during unroong of new uplifted regions such as the La
Cira-Infantas paleohighs.
4.2.4. Esmeraldas Formation
Seven sublitharenite to quartzarenite samples from the Esmeraldas Formation in the eastern limb of the syncline yield mean
compositions of Qm78-F4-Lt18 and Lm57-Lv9-Ls32. Although similar
to the upper La Paz Formation, the Esmeraldas shows a slight
increase in both plagioclase and potassium feldspar.
Nine medium-grained samples of litharenite were collected
from the Esmeraldas Formation in the western limb. The samples
dene a mean composition of Qm62-F9-Lt29 with an absence of
potassium feldspar and a greater proportion of lithic fragments
than the Esmeraldas samples from the eastern limb. Lithic
proportions are, however, largely similar between the two limbs
with a mean composition of Lm56-Lv8-Ls36.
4.2.4.1. Interpretations. Esmeraldas sandstone compositions in the
eastern limb indicate a source area with similar composition to that
of the upper La Paz Formation. However, the increase in potassium

259

feldspar may indicate sedimentation from crystalline basement


rocks of the Eastern Cordillera (e.g., Floresta and Santander Massifs;
Horton et al., 2010b). This change could also be due to variations in
the amount of preferential diagenetic elimination of feldspar
grains. A variation in composition between Esmeraldas samples in
the eastern and western limb suggests some degree of different
sediment sources, or downstream/lateral variations between the
two limbs.
5. Discussion
New results on Paleogene deposystems, sediment dispersal
patterns, and sandstone compositions reveal shifts in sediment
source areas for the Middle Magdalena Valley Basin. The implications of these shifts, as recorded at stratigraphic type localities in
the Nuevo Mundo Syncline, help delimit the temporal and spatial
evolution of structures within the basin and along the bounding
western and eastern basin margins (Central Cordillera and Eastern
Cordillera, respectively). Moreover, the activation of new
fold-thrust structures that helped generate the Eastern Cordillera
also succeeded in isolating the Middle Magdalena Valley Basin and
completing its conversion from a foreland basin to hinterland basin.
Sedimentological and provenance data from the Lisama
Formation indicate a mid-Paleocene shift in sediment sources
during a transition from marine deltaic/coastal to nonmarine

Fig. 10. Highly schematic cross sections depicting Paleogene basin evolution in the northern Middle Magdalena Valley, with the distribution of sediment sources reconstructed on
the basis of sedimentological, paleocurrent, and sandstone point-count data. Large arrows show generalized sediment dispersal patterns. Small arrows show fault displacement
with dashed lines indicating earliest possible deformation. (A) Earlyemiddle Paleocene sedimentation derived from the Guyana Shield. (B) Late Paleoceneeearly Eocene deposition
of sediment derived from initial topographic growth of the Central Cordillera. (C) Middleelate Eocene deposition of proximal sediment derived from uplift of La Cira-Infantas
paleohighs, which partially shuts off sediment delivery from Central Cordillera to the Nuevo Mundo Syncline. (D) Late Eoceneeearly Oligocene accumulation of sediment
largely derived from uplift of the Eastern Cordillera, which becomes the dominant sediment source to the Middle Magdalena Valley Basin.

260

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

conditions in the Middle Magdalena Valley Basin. Although


north-directed paleocurrents for the lowermost Lisama Formation
suggest broadly axial transport, the dominance of Proterozoic-age
zircon grains (Nie et al., 2010) implicate the eastern craton
(Guyana Shield) as the principal source region during the early
Paleocene (Fig. 10A). The possibility of Proterozoic zircons recycled
from basin ll in the west can be ruled out effectively by the lack of
Jurassic-Cretaceous zircons that would be expected from
magmatic-arc rocks of the Central Cordillera (Nie et al., 2010). Our
results reinforce this point, as there are no strong compositional
signatures of the considerable igneous detritus that would be
expected from erosion of the Central Cordillera.
For the late Paleocene, sedimentary indicators in the upper
Lisama Formation show an abrupt switch to east-directed paleocurrents, a higher proportion of volcanic lithic fragments, and
coarser deposits of principally subaerial delta plain systems. The
volcanic lithic fragments (felsitic, lathwork, and volcaniclastic
grains) are considered indicative of the magmatic-arc source areas
composing the bulk of the Central Cordillera (Fig. 10B). The Central
Cordillera contains both intrusive and extrusive magmatic belts
(Aspden and McCourt, 1986), including a volcaniclastic component
in the San Lucas Range of the northeastern Central Cordillera
(Kammer and Snchez, 2006) (Fig. 1). Detrital zircon UePb ages are
further consistent with a late Paleocene inux of Upper JurassiceLower Cretaceous (150e100 Ma) volcanic rocks from the Central
Cordillera into the Middle Magdalena Valley Basin (Nie et al., 2010).
Exhumation of the Central Cordillera at this time may be linked to
possible early displacement on the Cantagallo Thrust Fault or
related fold-thrust structures along the western margin of the
Magdalena Valley (Fig. 10B). Upper Paleocene to middle Eocene
strata (upper Lisama and lower La Paz Formations) show a progradational trend from a deltaic to distal uvial-plain setting
dominated by meandering channel to overbank systems. However,
consistent east-directed paleocurrents, sandstone compositions
rich in volcanic lithic fragments, and a continued presence of
arc-derived Mesozoic age zircons (Nie et al., 2010) point to the
Central Cordillera as the dominant source (Fig. 10B).
By the middleelate Eocene, the appearance of more quartzose
compositions in braided stream to uvial megafan facies of the
upper La Paz Formation suggests the introduction of a moreproximal source of sediment. This change is also reected in
sandstone compositions wherein upper La Paz samples show
signicant decreases in the fraction of total lithic fragments and
relative proportion of volcanic lithic fragments. Although paleocurrents continue to be east-directed, UePb geochronological
results also show a shift to much older, Grenville-age detrital
zircons (Nie et al., 2010). We attribute this shift to the initiation of
uplift of a new, proximal source area near the western ank of the
Nuevo Mundo Syncline. Because the zircon UePb ages and sediment compositions are similar to those of the Eastern Cordillera, it
seems that this new western source contained a bulk-rock geologic
column comparable to that of the Eastern Cordillera. We propose
that this source was most likely associated with mainly Eocene
uplift and unroong of rocks exposed in the La Cira-Infantas paleohighs, which may have formed a topographic barrier, effectively
shutting off sediment contributions from the Central Cordillera to
the Nuevo Mundo Syncline (Fig. 10C). In this interpretation, a critical observation is the westward thinning of the La Paz Formation in
the Nuevo Mundo Syncline (Fig. 3), which suggests that the upper
La Paz was likely deposited atop a growing structure. The precise
structural geometries are poorly known, and the growing structure
could have been linked at depth either westward with the La
Cira-Infantas paleohighs or eastward with the Lisama Anticline and
La Salina Thrust (Figs. 2B and 10C). Regardless of the exact linkage
and principal vergence direction (east- or west-directed faulting),

the provenance and stratigraphic patterns lead us to speculate that


upper La Paz deposition during the middleelate Eocene reects
proximal accumulation in the eastward-thickening wedge-top
depozone of a foreland basin associated with a complex thrust front
involving several buried structures (Fig. 10C). Upsection, continued
westward thinning of the Esmeraldas Formation is consistent with
progressive westward onlap onto the growing structures (Gmez
et al., 2003, 2005b) as sediment derived from west-owing rivers
accumulated against, and eventually on top of the beveled La
Cira-Infantas paleohighs.
Upper Eocene-lower Oligocene deposits of the lower Esmeraldas Formation recorded another major adjustment in paleoow,
from east-directed to west-directed. Lithofacies assemblages indicate a change toward more-distal sedimentation consisting of
lower gradient, meandering uvial deposits similar to those of the
lower La Paz Formation. Sandstone compositions and zircon UePb
ages are congruent with those of the underlying, middleeupper
Eocene strata. These factors lead us to interpret initial shorteningrelated exhumation of the western ank of the Eastern Cordillera
by latest Eocene time (Fig. 10D), a pattern consistent with subsurface growth strata reported along the La Salina Thrust and Nuevo
Mundo Syncline, the effective Magdalena Valley-Eastern Cordillera
boundary (Gmez et al., 2003, 2005b). Through the remainder of
the Paleogene, the principally recycled sedimentary materials from
the extensive Cretaceous succession across the Eastern Cordillera
continued to be the major source of sediment to the Nuevo Mundo
Syncline and Middle Magdalena Valley Basin.
The documented Paleogene transition from foreland to hinterland basin evolution underscores several points regarding the
Magdalena Valley and basin evolution in inter-Andean settings.
First, growth of the Eastern Cordillera as a topographic barrier to
foreland-directed sediment dispersal systems likely promoted
signicantly greater accumulation than in a conventional foreland
basin. Protracted subsidence in the eastern Middle Magdalena
Valley Basin has spanned >60 Myr, with accumulation of >10 km of
basin ll. The considerable duration and magnitude of accumulation of the Magdalena Valley matches or exceeds that of the Llanos
Basin in the modern foreland. Second, accommodation mechanisms in the Magdalena Valley showed considerable variation
through time. Although latest CretaceouseEocene accumulation is
largely the product of exural loading by tectonic thickening in the
growing Central Cordillera, subsidence driven by sediment loading
dominates the middle to late Cenozoic history of the Middle
Magdalena Valley Basin (Gmez et al., 2005b; Parra et al., 2009a).
We consider trapping or ponding of sediment on the hinterland
side of the Eastern Cordillera as a primary control on the long
duration and large scale of accumulation in the Middle Magdalena
Valley Basin as well as other Andean hinterland basins such as the
Altiplano-Puna system in the central Andes (e.g., Horton et al.,
2002; Leier et al., 2010; Murray et al., 2010; Horton, in press).
6. Conclusions
Paleogene siliciclastic deposits of the Nuevo Mundo Syncline
contain sedimentary lithofacies associations representing a transition from marginal marine to mixed meandering and braided
uvial systems in the Middle Magdalena Valley of Colombia.
Provenance signatures and considerations of structural history
further indicate Cenozoic sediment accumulation in a subsiding
region that evolved from a distal to proximal foreland basin and
then to a hinterland basin conguration. A several-phase reconstruction of clastic deposystems and provenance identies the
major sedimentation pathways and evolving sediment source areas
that can be linked to activity along various fold-thrust structures
during shortening in the northern Andes.

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

(1) Paleocene deltaic/coastal depositional environments were


replaced by an earlyemiddle Eocene uvial-plain setting
with principally meandering uvial systems providing arcderived sediment from the growing Central Cordillera in
the west. Following craton-derived sedimentation in the
latest Cretaceous to early Paleocene (Fig. 10A), the
mid-Paleocene record of the Lisama Formation shows
a switch from north- to east-directed paleocurrents and
a sharp increase in the volcanic lithic fraction of sandstone
compositions. We attribute these depositional and compositional changes to initial shortening-induced uplift of the
Central Cordillera, with potential early motion on the
east-directed Cantagallo Thrust Fault (Fig. 10B).
(2) By the middleelate Eocene, there was a shift to higher energy
depositional processes, as reected by braided uvial systems
(including possible uvial megafans) in the upper La Paz
Formation. Sandstone compositions also became more
quartzose at this time, showing a signicant decrease in total
lithic fragments and volcanic lithic fragments in particular. We
interpret this shift to represent sediment inux from a newly
uplifted proximal structure along the western ank of the
Nuevo Mundo Syncline. This middleelate Eocene sediment
was most likely derived from the adjacent La Cira-Infantas
paleohighs (presently buried in the Magdalena Valley subsurface), which may have acted as a partial topographic barrier
preventing eastward delivery of sediment originating in the
Central Cordillera (Fig. 10C).
(3) Beginning in the late Eocene, facies associations for the
Esmeraldas Formation suggest the elimination of a proximal
sediment source and return to meandering uvial conditions.
However, latest Eocene paleocurrents show a reversal from the
earlier east-directed transport to principally west-directed
paleoow. The record of ner grained, more-distal accumulation may imply a lowering of the gradient toward the Nuevo
Mundo Syncline, or greater distance to an emerging source
during early shortening-related growth of the Eastern Cordillera (Fig. 10D). This history indicates new topographic growth
to the east, a pattern consistent with the absence of arc-derived
zircons of Central Cordilleran provenance in the Oligocene and
younger succession. Moreover, deformation along the eastern
margin of the basin suggests the Middle Magdalena Valley
Basin was undergoing conversion from an initial foreland basin
to subsequent hinterland basin that persisted from the Oligocene to present.
In summary, we suggest that a hinterland basin overprinting or
succeeding a former foreland basin represents an underappreciated
but relatively common mode of basin evolution in the Andes. In the
eastern Middle Magdalena Valley Basin, >60 Myr of foreland and
hinterland accumulation has generated >10 km of basin ll. In this
case and in other inter-Andean settings, the growth of a topographic barrier in more-distal sectors of an early foreland basin
potentially acts as a sediment dam, forcing considerably greater
accumulation over longer periods without a direct need for
enhanced exural subsidence due to thrust loading.
Acknowledgments
Funding was provided by the Instituto Colombiano del Petrleo
(ICP), a division of Ecopetrol, and the Jackson School of Geosciences,
as part of a collaborative research agreement between ICP and the
University of Texas at Austin. The ICP-Ecopetrol project "Cronologia
de la deformacin en las Cuencas Subandinas" shared valuable
information and logistical support during the research. Additional
funding was provided by the University of Texas Graduate School

261

Diversity Mentoring Fellowship and the American Association of


Petroleum Geologists Foundation Grants-In-Aid of Research
program. We thank two anonymous reviewers and Associate Editor
Hermann Duque-Caro for helpful reviews that improved the
manuscript. Discussions with Junsheng Nie, Joel Saylor, Alejandro
Bande, Javier Sanchez, Richard Ketcham, Ron Steel, and Kitty Milliken helped clarify our interpretations and presentation. Technical
and eld assistance was provided by Julian Ramos, Jose Ricardo,
Jaime Corredor, Ana Milena Rangel, Eliseo Tesn, and Yimmy Corts.
References
Acosta, J., Velandia, F., Osorio, J., Lonergan, L., Mora, H., 2007. Strike-slip deformation
within the Colombian Andes. In: Ries, A.C., Butler, R.W.H., Graham, R.H. (Eds.),
Deformation of the Continental Crust: The legacy of Mike Coward. Geological
Society of London Special Publication, vol. 272, pp. 303e319.
Allen, J.R.L., 1965. A review of the origin and characteristics of recent alluvial
sediments. Sedimentology 5, 89e191.
Allmendinger, R.W., 1998. Inverse and forward numerical modeling of trishear
fault-propagation folds. Tectonics 17, 640e656.
Aspden, J.A., McCourt, W.J., 1986. Mesozoic oceanic terrane in the central Andes of
Colombia. Geology 14, 415e418.
Aspden, J.A., McCourt, W.J., Brook, M., 1987. Geometrical control of subductionrelated magmatism: the Mesozoic and Cenozoic plutonic history of Western
Colombia. Journal of the Geological Society 144, 893e905.
Bayona, G., Corts, M., Jaramillo, C., Ojeda, G., Aristizabal, J.J., Reyes-Harker, A., 2008.
An integrated analysis of an orogenesedimentary basin pair: latest CretaceouseCenozoic evolution of the linked Eastern Cordillera orogen and the
Llanos foreland basin of Colombia. Geological Society of America Bulletin 120,
1171e1197.
Bhattacharya, J.P., Giosan, L., 2003. Wave-inuenced deltas: geomorphological
implications for facies reconstruction. Sedimentology 50, 187e210.
Butler, K., Schamel, S., 1988. Structure along the eastern margin of the Central
Cordillera, Upper Magdalena Valley, Colombia. Journal of South American Earth
Sciences 1, 109e120.
Caballero, V., Evolucin tectono-sedimentaria del sinclinal de Nuevo Mundo, cuenca
sedimentaria Valle Medio del Magdalena Colombia, durante el OligoceneMioceno [M.S. thesis]: Bucaramanga, Colombia, Universidad Industrial de
Santander, 149 p.
Caballero, V., Parra, M., Mora, A., 2010. Levantamiento de la Cordillera Oriental de
Colombia durante el Eocene tardieOligoceno temprano: Proveniencia sedimentaria en el sinclinal de Nuevo Mundo, cuenca Valle Medio del Magdalena.
Boletn de Geologia 32, 45e77.
Colletta, B., 1990. Tectonic style and crustal structure of the Eastern Cordillera
(Colombia) from a balanced cross-section. In: Letouzey, J. (Ed.), Petroleum and
Tectonics in Mobile Belts. Editions Technip, Paris, pp. 81e100.
Cooper, M.A., Addison, F.T., Alvarez, R., Coral, M., Graham, R.H., Hayward, A.B.,
Howe, S., Martinez, J., Naar, J., Peas, R., Pulham, A.J., Taborda, A., 1995. Basin
development and tectonic history of the Llanos basin, eastern Cordillera, and
middle Magdalena valley, Colombia. American Association of Petroleum Geologists Bulletin 79, 1421e1443.
Cordani, U.G., Sato, K., Teixeira, W., Tassinari, C.C.G., Basei, M.A.S., 2000. Crustal
evolution of the South American platform. In: Cordani, U.G., Miliani, E.J., Thomaz-Filho, A., Campos, D.A. (Eds.), Tectonic Evolution of South America. 31st
International Geological Congress, Rio de Janeiro, Brazil, pp. 19e40.
Corredor, F., 2003. Eastward extent of the late Eocene-early Oligocene onset of
deformation across the northern Andes: constraints from the northern portion
of the Eastern Cordillera fold belt, Colombia. Journal of South American Earth
Sciences 16, 445e457.
Corts, M., Angelier, J., Colletta, B., 2005. Paleostress evolution of the northern
Andes (Eastern Cordillera of Colombia): implications on plate kinematics of the
south Caribbean region. Tectonics 24. doi:10.1029/2003TC001551 TC1008.
DeCelles, P.G., Cavazza, W., 1999. A comparison of uvial megafans in the Cordilleran (Upper Cretaceous) and modern Himalayan foreland basin systems.
Geological Society of America Bulletin 111, 1315e1334.
DeCelles, P.G., Langford, R.P., Schwartz, R.K., 1983. Two new methods of paleocurrent determination from trough cross-stratication. Journal of Sedimentary
Research 53, 629e642.
Dengo, C.A., Covey, M.C., 1993. Structure of the Eastern Cordillera of Colombia:
implications for trap styles and regional tectonics. American Association of
Petroleum Geologists Bulletin 77, 1315e1337.
Dickinson, W.R., 1970. Interpreting detrital modes of graywacke and arkose. Journal
of Sedimentary Petrology 40, 695e707.
Folk, R.L., 1980. Petrology of Sedimentary Rocks. Hemphill Publishing Company,
Austin, TX, 184 pp.
Gmez, E., 2001, Tectonic controls on the Late Cretaceous to Cenozoic sedimentary
ll of the Middle Magdalena Valley Basin, Eastern Cordillera and Llanos Basin,
Colombia [Ph.D. thesis]: Ithaca, New York, Cornell University, 619 p.
Gmez, E., Jordan, T.E., Allmendinger, R.W., Cardozo, N., 2005a. Development of the
Colombian foreland-basin system as a consequence of diachronous exhumation
of the northern Andes. Geological Society of America Bulletin 117, 1272e1292.

262

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263

Gmez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., Kelley, S., 2005b. Syntectonic Cenozoic sedimentation in the northern middle Magdalena valley basin of
Colombia and implications for exhumation of the northern Andes. Geological
Society of America Bulletin 117, 547e569.
Gmez, E., Jordan, T.E., Allmendinger, R.W., Hegarty, K., Kelley, S., Heizler, M., 2003.
Controls on architecture of the late Cretaceous to Cenozoic southern middle
Magdalena valley basin, Colombia. Geological Society of America Bulletin 115,
131e147.
Galloway, W.E., 1976. Sediments and stratigraphic framework of the copper river
fan-delta, Alaska. Journal of Sedimentary Research 46, 726e737.
Gazzi, P., 1966. I minerali pesanti nei ysch arenacei fra Monte Ramaceto e Monte
Molinatico (Appennino settentrionale). Mineralogica et Petrographica Acta 11,
197e212.
Hampton, B.A., Horton, B.K., 2007. Sheetow uvial processes in a rapidly subsiding
basin, Altiplano plateau, Bolivia. Sedimentology 54, 1121e1147.
Hopping, C.A., 1967. Palynology and the oil industry. Review of Palaeobotany and
Palynology 2, 23e48.
Horton, B.K., 2005. Revised deformation history of the central Andes: inferences
from Cenozoic foredeep and intermontane basins of the Eastern Cordillera,
Bolivia. Tectonics 24. doi:10.1029/2003TC001619 TC3011.
Horton, B.K., DeCelles, P.G., 1997. The modern foreland basin system adjacent to the
Central Andes. Geology 25, 895e898.
Horton, B.K., DeCelles, P.G., 2001. Modern and ancient uvial megafans in the
foreland basin system of the central Andes, southern Bolivia: implications for
drainage network evolution in fold-thrust belts. Basin Research 13, 43e63.
Horton, B.K., Hampton, B.A., LaReau, B.N., Baldelln, E., 2002. Tertiary provenance
history of the northern and central Altiplano (central Andes, Bolivia): a detrital
record of plateau-margin tectonics. Journal of Sedimentary Research 72,
711e726.
Horton, B.K., Parra, M., Saylor, J.E., Nie, J., Mora, A., Torres, V., Stockli, D.F.,
Strecker, M.R., 2010a. Resolving uplift of the northern Andes using detrital
zircon age signatures. GSA Today 20, 4e9,. doi:10.1130/GSATG76A.1.
Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., Stockli, D.F.,
2010b. Linking sedimentation in the northern Andes to basement conguration,
Mesozoic extension, and Cenozoic shortening: evidence from detrital zircon
UePb ages in the Eastern Cordillera of Colombia. Geological Society of America
Bulletin 122, 1423e1442.
Horton, B.K., Cenozoic evolution of hinterland basins in the Andes and Tibet, in
Busby, C., and Azor, A., eds., Tectonics of Sedimentary Basins: Recent Advances:
Wiley-Blackwell, in press.
Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., Sares, S.W., 1984. The
effect of grain size on detrital modes: a test of the Gazzi-Dickinson pointcounting method. Journal of Sedimentary Petrology 54, 103e116.
Jordan, T.E., 1995. Retroarc foreland and related basins. In: Busby, C.J., Ingersoll, R.V.
(Eds.), Tectonics of Sedimentary Basins. Blackwell Science, Cambridge, Massachusetts, pp. 331e362.
Jordan, T.E., Alonso, R.N., 1987. Cenozoic stratigraphy and basin tectonics of the
Andes Mountains, 20 e28 south latitude. American Association of Petroleum
Geologists Bulletin 71, 49e64.
Julivert, M., 1970. Cover and basement tectonics in the Cordillera oriental of
Colombia, South America, and a comparison with some other folded chains.
Geological Society of America Bulletin 81, 3623e3646.
Kammer, A., Snchez, J., 2006. Early Jurassic rift structures associated with the
Soapaga and Boyac faults of the Eastern Cordillera, Colombia: sedimentological inferences and regional implications. Journal of South American Earth
Sciences 21, 412e422.
Kellogg, J.N., Vega, V., 1995. Tectonic development of Panama, Costa Rica, and the
Colombian Andes; constraints from global positioning system geodetic studies
and gravity. Geological Society of America Special Paper 295, 75e90.
Leier, A.L., McQuarrie, N., Horton, B.K., Gehrels, G.E., 2010. Upper Oligocene
conglomerates of the Altiplano, central Andes: the record of deposition and
deformation along the margin of a hinterland basin. Journal of Sedimentary
Research 80, 750e762.
Linke, P., Lutze, G., 1993. Microhabitat preferences of benthic foraminiferaea static
concept or a dynamic adaptation to optimize food acquisition. Marine Micropaleontology 20, 215e234.
Makaske, B., 2001. Anastomosing rivers: a review of their classication, origin and
sedimentary products. Earth-Science Reviews 53, 149e196.
Marocco, R., Lavenu, A., Baudino, R., 1995. Intermontane late PaleogeneeNeogene
basins of the Andes of Ecuador and Peru: sedimentologic and tectonic characteristics. In: Tankard, A.J., Surez, R., Welsink, H.J. (Eds.), Petroleum Basins of
South America. American Association of Petroleum Geologists Memoir, vol. 62,
pp. 597e613.
Marsaglia, K.M., Ingersoll, R.V., 1992. Compositional trends in arc-related, deepmarine sand and sandstone: a reassessment of magmatic-arc provenance.
Geological Society of America Bulletin 104, 1637e1649.
McCourt, W.J., Aspden, J.A., Brook, M., 1984. New geological and geochronological
data from the Colombian Andes: continental growth by multiple accretion.
Journal of the Geological Society of London 141, 831e845.
Miall, A.D., 1977. A review of the braided-river depositional environment. EarthScience Reviews 13, 1e62.
Miall, A.D., 1985. Architectural-element analysis: a new method of facies analysis
applied to uvial deposits. Earth-Science Reviews 22, 261e308.
Miall, A.D., 1996. The Geology of Fluvial Deposits: Sedimentary Facies, Basin Analysis, and Petroleum Geology. Springer-Verlag, Berlin, 582 pp.

Milliken, K.L., 1988. Loss of provenance information through subsurface diagenesis


in PlioePleistocene sandstones, northern Gulf of Mexico. Journal of Sedimentary Petrology 58, 992e1002.
Milliken, K.L., 1992. Chemical behavior of detrital feldspars in mudrocks versus
sandstones, Frio Formation (Oligocene), south Texas. Journal of Sedimentary
Petrology 62, 790e801.
Milliken, K.L., McBride, E.F., Land, L.S., 1989. Numerical assessment of dissolution
versus replacement in the subsurface destruction of detrital feldspars, Oligocene
Frio Formation, south Texas. Journal of Sedimentary Petrology 59, 740e757.
Montes, C., Hatcher Jr., R.D., Restrepo-Pace, P.A., 2005. Tectonic reconstruction of the
northern Andean blocks: oblique convergence and rotations derived from the
kinematics of the PiedraseGirardot area, Colombia. Tectonophysics 399, 221e250.
Mora, A., Parra, M., Strecker, M.R., Kammer, A., Dimat, C., Rodrguez, F., 2006.
Cenozoic contractional reactivation of Mesozoic extensional structures in the
Eastern Cordillera of Colombia. Tectonics 25. doi:10.1029/2005TC001854 2010.
Mora, A., Parra, M., Strecker, M.R., Sobel, E.R., Hooghiemstra, H., Torres, V.,
Jaramillo, J.V., 2008. Climatic forcing of asymmetric orogenic evolution in the
Eastern Cordillera of Colombia. Geolical Society of America Bulletin 120,
930e949.
Mora, A., Horton, B.K., Mesa, A., Rubiano, J., Ketcham, R.A., Parra, M., Blanco, V.,
Garcia, D., Stockli, D.F., 2010. Migration of Cenozoic deformation in the Eastern
Cordillera of Colombia interpreted from ssion track results and structural
relationships: implications for petroleum systems. American Association of
Petroleum Geologists Bulletin 94, 1543e1580.
Morales, L.G., Podesta, D.J., Hateld, W.C., Tanner, H., Jones, S.H., Barker, M.H.S.,
ODonoghue, D.J., Mohler, C.E., Dubois, E.P., Jacobs, C., Goss, C.R., 1958. General
geology and oil occurrences of middle Magdalena valley, Colombia. In:
Weeks, L.G. (Ed.), Habitat of Oil Symposium. American Association of Petroleum
Geologists, pp. 641e695.
Murray, B.P., Horton, B.K., Matos, R., Heizler, M.T., 2010. Oligocene-Miocene basin
evolution in the northern Altiplano, Bolivia: implications for evolution of the
central Andean backthrust belt and high plateau. Geological Society of America
Bulletin 122, 1443e1462.
Nanson, G.C., Knighton, A.D., 1996. Anabranching rivers: their cause, character and
classication. Earth Surface Processes and Landforms 21, 217e239.
Nie, J., Horton, B.K., Mora, A., Saylor, J.E., Housh, T.B., Rubiano, J., Naranjo, J., 2010.
Tracking exhumation of Andean ranges bounding the middle Magdalena valley
basin, Colombia. Geology 38, 451e454.
Orton, G.J., Reading, H.G., 1993. Variability of deltaic processes in terms of sediment
supply, with particular emphasis on grain size. Sedimentology 40, 475e512.
Pardo-Trujillo, A., 2004, Paleocene-Eocene palynology and palynofacies from
northeastern Colombia and western Venezuela [Ph.D. thesis]: Lige, Belgium,
Universit de Lige, 103 p.
Pardo-Trujillo, A., Jaramillo, C., 2002. New palynostratigraphical data of NW south
America. PaleoceneeEocene of the middle Magdalena valley, Colombia. International Journal of Tropical Geology, Geography and Ecology 26 (1), 1e10.
Pardo-Trujillo, A., Jaramillo, C.A., Oboh-Ikuenobe, F.E., 2003. Paleogene palynostratigraphy of the eastern middle Magdalena Valley, Colombia. Palynology 27, 155e178.
Parra, M., Mora, A., Jaramillo, C., Strecker, M.R., Sobel, E.R., Quiroz, L., Rueda, M.,
Torres, V., 2009a. Orogenic wedge advance in the northern Andes: evidence
from the Oligocene-Miocene sedimentary record of the Medina basin, eastern
Cordillera, Colombia. Geological Society of America Bulletin 121, 780e800.
Parra, M., Mora, A., Sobel, E.R., Strecker, M.R., Gonzlez, R., 2009b. Episodic orogenic
front migration in the northern Andes: constraints from low-temperature
thermochronology in the Eastern Cordillera, Colombia. Tectonics 28.
doi:10.1029/2008TC002423 TC4004.
Pennington, W.D., 1981. Subduction of the eastern Panama basin and seismotectonics of northwestern South America. Journal of Geophysical Research 86,
10753e10770.
Pilsbry, H.A., and Olsson, A.A., 1935, Tertiary fresh-water mollusks of the Magdalena
Embayment, Colombia; with Tertiary stratigraphy of the middle Magdalena
Valley by O. C. Wheeler: Proceedings of the Academy of Natural Sciences,
Philadelphia, v. 87, pp. 7e39.
Ramn, J.C., 1998, Sequence stratigraphic framework of Tertiary strata and oil
geochemical evaluation, Middle Magdalena Basin, Colombia [Ph.D. thesis]:
Golden, Colorado, Colorado School of Mines, 270 p.
Ramirez, R.E., 1988, Stratigraphy of the Tertiary of the Middle Magdalena Basin
(Colombia), central and northern parts [M.S. thesis]: Austin, Texas, The
University of Texas, 199 p.
Ramon, J.C., Rosero, A., 2006. Multiphase structural evolution of the western margin
of the Girardot subbasin, upper Magdalena valley, Colombia. Journal of South
American Earth Sciences 21, 493e509.
Restrepo-Pace, P.A., Colmenares, F., Higuera, C., Mayorga, M., 2004. A fold-andthrust belt along the western ank of the Eastern Cordillera of Colombiadstyle,
kinematics, and timing constraints derived from seismic data and detailed
surface mapping. In: McClay, K.R. (Ed.), Thrust Tectonics and Hydrocarbon
Systems. American Association of Petroleum Geologists Memoir, vol. 82, pp.
598e613.
Rincn, D.A., Arenas, J.E., Cuartas, C.H., Crdenas, A.L., Molinares, C.E., Caicedo, C.,
Jaramillo, C., 2007. Eocene-Pliocene planktonic foraminifera biostratigraphy
from the continental margin of the southwest Caribbean. Stratigraphy 4,
261e311.
Rolon, L.F., 2004, Structural geometry of the Jura-Cretaceous rift of the Middle
Magdalena Valley Basin e Colombia [M.S. thesis]: Morgantown, West Virginia,
West Virginia University, 63 p.

C.J. Moreno et al. / Journal of South American Earth Sciences 32 (2011) 246e263
Sarmiento-Rojas, L., Van Wess, J., Cloetingh, S., 2006. Mesozoic transtensional basin
history of the Eastern Cordillera, Colombian Andes: inferences from tectonic
models. Journal of South American Earth Sciences 21, 383e411.
Saylor, J.E., Horton, B.K., Nie, J., Corredor, J., Mora, A., 2011. Evaluating foreland basin
partitioning in the northern Andes using Cenozoic ll of the Floresta basin,
Eastern Cordillera, Colombia. Basin Research 23, 377e402. doi:10.1111/j.13652117.2010.00493.x.
Schamel, S., 1991. Middle and upper Magdalena basins, Colombia. In: Biddle, K.T.
(Ed.), Active Margin Basins. American Association of Petroleum Geologists
Memoir, vol. 52, pp. 283e301.
Smoot, J.P., 1991. Sedimentary facies and depositional environments of early
Mesozoic Newark supergroup basins, eastern North America. Palaeogeography,
Palaeoclimatology, Palaeoecology 84, 369e423.
Surez, M.A., 1997, Facies analysis of the Upper Eocene La Paz Formation, and
regional evaluation of the post-middle Eocene stratigraphy, northern Middle
Magdalena Valley Basin, Colombia [M.S. thesis]: Boulder, Colorado, University
of Colorado, 88 p.
Suppe, J., Chou, G.T., Hook, S.C., 1992. Rates of folding and faulting determined from
growth strata. In: McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall,
London, pp. 105e122.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
Rivera, C., 2000. Geodynamics of the northern Andes: subductions and intracontinental deformation (Colombia). Tectonics 19, 787e813.

263

Teixeira, W., Tassinari, C.C.G., Cordani, U.G., Kawashita, K., 1989. A review of the
geochronology of the Amazonian Craton: tectonic implications. Precambrian
Research 42, 213e227.
Tye, R.S., Bhattacharya, J.P., Lorsong, J.A., 1999. Geology and stratigraphy of uviodeltaic deposits in the Ivishak formation: applications for development of
Prudhoe Bay eld Alaska. American Association of Petroleum Geologists
Bulletin 83, 1588e1623.
Uba, C.E., Heubeck, C., Hulka, C., 2005. Facies analysis and basin architecture of the
Neogene Subandean synorogenic wedge, southern Bolivia. Sedimentary
Geology 180, 91e123.
Van Houten, F.B., Travis, R.B., 1968. Cenozoic deposits, upper Magdalena valley,
Colombia. American Association of Petroleum Geologists Bulletin 52,
675e702.
van der Hilst, R., Mann, P., 1994. Tectonic implications of tomographic images of
subducted lithosphere beneath northwestern South America. Geology 22,
451e454.
Van Houten, F.B., 1976. Late Cenozoic volcaniclastic deposits, Andean foredeep,
Colombia. Geological Society of America Bulletin 87, 481e495.
Villamil, T., 1999. Campanian-Miocene tectonostratigraphy, depocenter evolution
and basin development of Colombia and western Venezuela. Palaeogeography,
Palaeoclimatology, Palaeoecology 153, 239e275.
Willis, B.J., Gabel, S., 2001. Sharp-based, tide-dominated deltas of the Sego sandstone, Book cliffs, Utah, USA. Sedimentology 48, 479e506.

S-ar putea să vă placă și